154 kriegsfeld et al

download 154 kriegsfeld et al

of 18

Transcript of 154 kriegsfeld et al

  • 7/31/2019 154 kriegsfeld et al

    1/18

    Review

    The regulation of neuroendocrine function: Timing is everything

    Lance J. Kriegsfeld a,, Rae Silverb,c,d

    a Department of Psychology and Helen Wills Neuroscience Institute, 3210 Tolman H all, #1650, University of California, Berkeley, CA 94720-1650, USAb Department of Psychology, Barnard College, New York, N Y 10027, USA

    c Department of Psychology, Columbia University, New York, NY 10027, USAd Department of Anatomy and Cell Biology, College of Physicians and Surgeons, New York, NY 10032, USA

    Received 24 September 2005; revised 6 December 2005; accepted 8 December 2005

    Available online 21 February 2006

    Abstract

    Hormone secretion is highly organized temporally, achieving optimal biological functioning and health. The master clock located in the

    suprachiasmatic nucleus (SCN) of the hypothalamus coordinates the timing of circadian rhythms, including daily control of hormone secretion. In

    the brain, the SCN drives hormone secretion. In some instances, SCN neurons make direct synaptic connections with neurosecretory neurons. In

    other instances, SCN signals set the phase of clock genes that regulate circadian function at the cellular level within neurosecretory cells. The

    protein products of these clock genes can also exert direct transcriptional control over neuroendocrine releasing factors. Clock genes and proteins

    are also expressed in peripheral endocrine organs providing additional modes of temporal control. Finally, the SCN signals endocrine glands via

    the autonomic nervous system, allowing for rapid regulation via multisynaptic pathways. Thus, the circadian system achieves temporal regulation

    of endocrine function by a combination of genetic, cellular, and neural regulatory mechanisms to ensure that each response occurs in its correct

    temporal niche. The availability of tools to assess the phase of molecular/cellular clocks and of powerful tract tracing methods to assess

    connections between clock cells and their targets provides an opportunity to examine circadian-controlled aspects of neurosecretion, in the

    search for general principles by which the endocrine system is organized.

    2005 Elsevier Inc. All rights reserved.

    Keywords: Circadian; Diurnal; Endocrinel; Neurosecretion; Clock genes; Suprachiasmatic

    Contents

    Circadian aspects of reproduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 558

    Circadian control of endocrine secretions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559

    Endocrine influences on the circadian system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559

    Identification of a brain clock: from tissue to gene. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560

    Circadian output and orchestration of endocrine function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560

    Diffusible signals controlling behavioral rhythms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560

    Neural control of neurosecretory factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 563

    Neural SCN output and estrus regulation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 563Direct and indirect transcriptional control as a clock output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 565

    Direct transcriptional control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 565

    Indirect transcriptional control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566

    System-level control and coordination of endocrine function. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 566

    Clocks in the neuroendocrine system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567

    The top of the hierarchy: neural SCN output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567

    Hormonal and neural communication to glands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 567

    Hormones and Behavior 49 (2006) 557574www.elsevier.com/locate/yhbeh

    Corresponding author. Fax: +1 510 642 5293.

    E-mail address: [email protected] (L.J. Kriegsfeld).

    0018-506X/$ - see front matter 2005 Elsevier Inc. All rights reserved.doi:10.1016/j.yhbeh.2005.12.011

    mailto:[email protected]:[email protected]://dx.doi.org/10.1016/j.yhbeh.2005.12.011http://dx.doi.org/10.1016/j.yhbeh.2005.12.011http://dx.doi.org/10.1016/j.yhbeh.2005.12.011mailto:[email protected]
  • 7/31/2019 154 kriegsfeld et al

    2/18

    Conclusions and perspectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 568

    Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569

    Circadian aspects of reproduction

    The importance of circadian (about a day) timing in

    hormone production/secretion has been known since the

    1950s when Everett and Sawyer determined that a

    stimulatory signal occurring during a narrow temporal

    window on the afternoon of proestrus is necessary for

    induction of ovulation later that night (Everett and Sawyer,

    1950). Such close temporal organization is important for

    successful reproduction, as numerous hormone-dependent

    behavioral and physiological processes must be coordinated.

    If optimal temporal relationships are disrupted, pronounced

    deficits in fertility can result. For example, ovulation,

    behavioral estrus, fertilization, and pregnancy maintenancerequire a specific temporal pattern of hormone secretion in

    spontaneous ovulators such as rats, hamsters, and mice

    (Blaustein et al., 1994; McEwen et al., 1987; Mong et al.,

    2003). Prior to behavioral estrus, rising levels of estrogen

    both trigger a precisely timed preovulatory surge in

    luteinizing hormone (LH) and stimulate the production of

    brain progesterone receptors in preparation for progesterone

    effects on neurons. The timing of progesterone receptor

    regulation relative to estrogen ensures that behavioral

    receptivity is coordinated with the time of ovulation,

    thereby increasing the likelihood of pregnancy (Hansen et

    al., 1979). Following ovulation, a prolactin surge is

    necessary to support the corpus luteum to maintainprogesterone secretion necessary for pregnancy and its

    maintenance (Egli et al., 2004). This sequence is under

    circadian control.

    In the present review, we summarize evidence indicating that

    the timing of endocrine secretions is coordinated by a brainclock located in the suprachiasmatic nucleus (SCN) of the

    hypothalamus (Fig. 1). Over the last decade, there have been

    substantial, rapid advances in our understanding of the circadian

    modulation of brain and peripheral organ activity. Current data

    indicate that a neural signal from the SCN is necessary for the

    circadian timing of hormone secretion, while a diffusible signal

    is sufficient to modulate non-endocrine events such as daily

    behavioral activities. The identification of core clock genes,

    clock-controlled genes (CCG), and their localization in

    neurosecretory cells (Kriegsfeld et al., 2003; Olcese et al.,

    2003), the pituitary gland (Shieh, 2003; Von Gall et al., 2002)

    and a number of peripheral endocrine glands (Bittman et al.,2003; Morse et al., 2003; Zylka et al., 1998) each provide new

    opportunities for evaluating the loci and mechanisms of

    temporal gating of hormone secretion.

    We review evidence that hormone secretion is regulated not

    only by the feedback loops long studied by endocrinologists but

    also by the SCN and SCN-derived temporal signals acting

    directly on neurosecretory cells, on the autonomic nervous

    system, and on clock genes and clock-controlled genes. To this

    end, we present an abbreviated introduction to the core negative

    feedback loop controlling cellular circadian clock function to

    provide a basis for understanding how time can be tracked

    within a cell and to set the foundation for understanding the

    possible role of clock genes in endocrine regulation. Fordetailed summaries of research on cellular/molecular clock

    genes and proteins, the reader is referred to the following

    reviews (Albrecht, 2004; Ashmore and Sehgal, 2003; Du and

    Fig. 1. The mammalian circadian clock is located in the suprachiasmatic nucleus (SCN) of the anterior hypothalamus. The SCN pictured here in this schematic is a

    coronal section through a rodent brain. The SCN is situated at the base of the brain of the brain directly above the optic chiasm (oc) and surrounding the third ventricle(V3). The sagittal schematic in the upper right corner depicts the approximate rostralcaudal location depicted in the coronal section.

    558 L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    3/18

    Tong, 2002; Duffield, 2003; Gachon et al., 2004; Glossop and

    Hardin, 2002; Green, 2003; Hastings et al., 2003; Liu, 2003;

    Okamura, 2003a,b; Okamura et al., 2002; Piggins, 2002;

    Roenneberg and Merrow, 2003; Schibler and Sassone-Corsi,

    2002; Schultz and Kay, 2003; Zordan et al., 2003).

    Circadian control of endocrine secretions

    Biological events typically exhibit marked, predictable

    cycles ranging in time from seconds to years. In the endocrine

    system, virtually every factor measured to date shows a

    circadian (endogenous) or diurnal (driven) rhythm. These

    alterations are achieved by modulation of pulse amplitude

    (i.e., amount of hormone released), pulse frequency (i.e., rate of

    hormone release), or by a combination of both of these

    processes. For example, studies in male rhesus macaques in

    which animals were sampled at 20-min intervals in an LD cycle

    revealed diurnal rhythms in luteinizing hormone (LH),

    testosterone, prolactin, and cortisol (Plant, 1981). Likewise,studies in rats, Syrian hamsters, and humans indicate circadian

    variation in gonadotropins and gonadal steroids around the

    onset of puberty (Andrews and Ojeda, 1981; Boyar et al., 1976;

    de la Iglesia et al., 1999; Jakacki et al., 1982; Smith and Stetson,

    1980). It has been suggested that diurnal variation in hormone

    concentrations may simply be modulated by sleep. However,

    sleep reversal (subjects sleep during the day rather than night)

    and sleep interruption do not affect the daily pattern of most

    hormones, confirming regulation by an endogenous clock

    independent of sleep (Desir et al., 1982; Kapen et al., 1974; Van

    Cauter and Refetoff, 1985), although interactions between sleep

    and the circadian system exist (Kriegsfeld et al., 2002a). The

    present review highlights the current understanding of circadiansystem regulation of endocrine rhythms via multiple means of

    modulation and proposes a novel mechanism of hierarchical

    control beginning with the master brain clock. For a complete

    description of daily hormone secretion patterns, the reader is

    referred to the following reviews (Gore, 1998; Hastings, 1991;

    Kriegsfeld et al., 2002a; Turek and Van Cauter, 1994).

    Endocrine influences on the circadian system

    Not only does the SCN regulate endocrine rhythms but

    hormones also feed back to the SCN, presumably to fine-

    tune the temporal pattern of endocrine secretion givencurrent conditions. For example, high-affinity melatonin

    receptors are localized to the SCN, and administration of

    melatonin can alter SCN phase (Dubocovich et al., 1996;

    Hastings et al., 1997; Lewy et al., 1992; Slotten et al., 1999;

    Vanecek and Watanabe, 1999). In addition, exogenous

    melatonin alters the phase of SCN electrical activity measured

    in a slice preparation in vitro in a manner predicted by the

    phase response curve (McArthur et al., 1991). The sensitivity

    of the SCN to melatonin may be a function of daily variation

    in the density of melatonin binding sites within the SCN

    (Schuster et al., 2001). In humans, melatonin administration

    causes phase delays during late night or early morning and

    phase advances in late morning to early afternoon (Lewy and

    Sack, 1997). This finding has significant implications for shift

    workers, jet lag, and the blind.

    Estrogen has pronounced effects on circadian activity

    rhythms, likely through both direct effects on the SCN and

    indirect mechanisms. In females, estrogen modulates both

    period and activity consolidation, suggesting actions on the

    circadian clock rather than transient effects on SCN targets.Cycling female hamsters and rats show a phase advance in

    locomotor activity on the day of estrous (scalloping), when

    estradiol levels are highest, and continuous administration of

    estradiol in silastic capsules shortens the free-running period of

    ovariectomized hamsters (Morin et al., 1977). When hamsters

    are maintained in constant light, the normally stable activity

    phase frequently splits into two activity components that

    stabilize approximately 12 h apart. Continuous administration

    of estradiol in silastic capsules to ovariectomized hamsters

    prevents these changes (Morin, 1980).

    Direct effects of estrogen on the circadian clock are

    suggested by the expression of and estrogen receptors inthe SCN across mammals, including humans (Gundlah et al.,

    2000; Kruijver and Swaab, 2002; Su et al., 2001). These

    receptors may be important for its normal development and

    synchronization to the environment (Abizaid et al., 2004;

    Gundlah et al., 2000). In humans, the presence of estrogen

    receptor expression, along with sex differences in SCN

    structure, suggests that estrogen may act on the SCN during

    development (Hofman et al., 1988, 1996; Kruijver and Swaab,

    2002). Indirect effects in estrogen on the circadian clock are

    indicated by studies in which simultaneous injection of

    anterograde and retrograde tract tracers into the SCN reveal

    that ER-expressing cells in the preoptic area, amygdala,

    BNST, and arcuate provide input to the SCN, but the SCN doesnot project directly to these ER-expressing cells (de la Iglesia

    et al., 1999).

    In males, testosterone also affects consolidation of locomotor

    activity rhythms. Extended exposure to short day lengths

    induces a decrease in testicular size and a decline in plasma

    testosterone concentrations in male hamsters (Ellis and Turek,

    1983). Following testicular regression (or after castration), there

    is an increase in lability of activity onset, an expansion of the

    daily activity duration, with a decrease in wheel revolutions per

    cycle; testosterone replacement prevents these changes (Morin

    and Cummings, 1981). Testosterone may act through SCN

    androgen receptors. To date, androgen receptors have beenidentified in the SCN of several species (Clancy et al., 1994;

    Fernandez-Guasti et al., 2000; Kashon et al., 1996; Michael and

    Rees, 1982; Rees and Michael, 1982). Alternatively, testoster-

    one may exert its effects through conversion to estradiol, which

    may act either directly on receptors in the SCN (e.g., Shughrue

    et al., 1997), or indirectly in ER-expressing cells in other brain

    areas that, in turn, communicate with the SCN (de la Iglesia et

    al., 1999). In rats, conversion of testosterone to estradiol may be

    important for the activity-stimulating effects of testosterone

    (Roy and Wade, 1975). Estradiol is nearly 100 times as effective

    as testosterone at increasing activity, while dihydrotesosterone

    (a non-aromatizable androgen) has no effect on wheel-running

    activity of rats (Roy and Wade, 1975). Taken together, these

    559L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    4/18

    findings suggest that the conversion of testosterone to estrogen,

    by aromatase, may be important for the effects of testosterone

    on circadian rhythms.

    Identification of a brain clock: from tissue to gene

    A highly localized brain clock in mammals was firstsuggested in 1972, with the demonstration that lesions ablating

    the SCN abolish circadian rhythmicity in adrenal corticoid

    secretion and locomotor behavior (Moore and Eichler, 1972;

    Stephan and Zucker, 1972). SCN-lesioned animals continue to

    show the full range of normal behaviors, but their temporal

    organization is lost and never recovers, irrespective of how early

    in development the lesions are performed (Mosko and Moore,

    1979). The initial conclusion that the SCN serves as a brain

    master clock has been confirmed in the subsequent 30 years by

    converging lines of research involving in vivo, ex vivo, and in

    vitro studies carried out in many different laboratories. For

    example, transplants of donor SCN tissue into the brains ofarrhythmic, SCN-lesioned hosts restore circadian rhythmicity in

    behavior (Lehman et al., 1987; Ralph et al., 1990). Importantly,

    rhythms are restored with the period of the donor SCN,

    indicating that the transplanted tissue does not act by restoring

    host brain function but that the clock is contained in the

    transplanted tissue. Further evidence that clock function is

    contained within the SCN comes from studies demonstrating

    that circadian rhythms in neural firing rate persist in isolated

    SCN tissue maintained in culture (Green and Gillette, 1982;

    Groos and Hendriks, 1982; Shibata et al., 1982). An excellent

    overview of these studies in historical perspective is available

    (Weaver, 1998).

    While the foregoing work demonstrated that the SCN tissueas a whole served as a clock, the finding that circadian

    rhythmicity is a property of individual SCN neurons set the

    stage for the next breakthrough. The demonstration that

    dispersed, cultured SCN cells exhibit circadian rhythms of

    electrical activity indicated that circadian timing is a cellular

    property rather than an emergent property of a neural network

    (Welsh et al., 1995). These studies allowed for the exploration

    and subsequent discovery of the cellular molecular machinery

    responsible for circadian function.

    Within a cell, circadian rhythms are produced by an

    autoregulatory transcriptional/translational negative feedback

    loop that takes approximately 24 h, whereas the generalmechanism for circadian oscillations at the cellular level is

    common among organisms, the components comprising the

    feedback loop differ. For the purpose of clarity, only the core

    mammalian feedback loop is described. To date, it is thought

    that two proteins, CLOCK and BMAL1, bind to one another

    and drive the transcription of messenger RNA (mRNA) of the

    Period (Per) and Cryptochrome (Cry) genes by binding to the

    E-box (CACGTG) domain on these gene promoters. Three

    Period (Per1, Per2, and Per3) and two cryptochrome genes

    (Cry1 and Cry2) have been identified. The mRNA for these

    genes is translated into PER and CRY proteins in the cytoplasm

    of the cell over the course of the day. Throughout the day, these

    proteins build up within the cytoplasm, and when they reach

    high enough levels, they form hetero- and homo-dimers. These

    newly formed dimers then feed back to the nucleus where they

    bind to the CLOCK:BMAL1 protein complex to turn off their

    own transcription (Fig. 2). Numerous otherclock genes and

    regulatory enzymes have been identified but will not be

    reviewed for the sake of brevity. Future studies on the specific

    genes and their interactions that result in circadian timekeepingat the cellular level will likely yield exciting new information on

    other regulatory elements and their interactions.

    Discovery of the genes regulating circadian rhythmicity led

    to breakthroughs identifying clock genes and their protein

    products in numerous sites, including extra-SCN brain loci and

    in the periphery (Abe et al., 2002; Balsalobre et al., 1998;

    Kriegsfeld et al., 2003; Yamazaki et al., 2000). These findings,

    in turn, led to questions about the unique nature of the master

    oscillator in the SCN, the functional significance of extra-SCN

    oscillators, and mechanisms of coordination of these widely

    dispersed clocks.

    To compare cellular mechanisms of clock gene expression inthe SCN and in the periphery, embryonic fibroblasts from wild-

    type and (behaviorally arrhythmic) Cry/ mice were used

    (Yagita et al., 2001). Clock properties of cell lines derived from

    peripheral cells of each strain were similar to those of the strain-

    specific SCN, supporting the conclusion of common core clock

    gene function in all tissues (Yagita et al., 2001). It remains

    controversial, however, whether peripheral oscillators are

    similar to those of the SCN in their ability to sustain endogenous

    rhythmicity for long durations (Balsalobre et al., 1998; Yoo et

    al., 2004).

    When a tissue, either SCN or peripheral, loses coherent

    rhythmicity, it is important to determine whether this is due to

    dampening of rhythms in individual cells or to loss ofsynchrony among a population of cells in the tissue. Use of

    Per1-luciferase transgenic animals indicates that rhythms in

    peripheral tissues damp then disappear over time due to

    uncoupling (desynchronization) among oscillators that retain

    their individual rhythms (Nagoshi et al., 2004; Welsh et al.,

    2004). Presumably peripheral clock cells normally get phase

    information (directly or indirectly) from the SCN to synchronize

    individual oscillators to each other. In this view, the SCN sets

    the phase of peripheral circadian clocks daily, coordinating the

    activity of tissues and organs of the body relative to one another,

    thereby maintaining homeostasis.

    Circadian output and orchestration of endocrine function

    Diffusible signals controlling behavioral rhythms

    Rhythmic electrical activity and oscillation of clock genes

    within the SCN neurons ultimately lead to rhythmicity in the

    whole organism. Compelling evidence for a diffusible output

    signal derives from neural tissue transplantations in which the

    SCN from a fetal donor is implanted into the third ventricle of

    an adult, SCN-lesioned host. As mentioned previously, these

    grafts restore activity-related behaviors such as locomotor,

    drinking, and gnawing rhythms (Lehman et al., 1987; Ralph et

    al., 1990; Silver et al., 1990). That a diffusible signal is

    560 L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    5/18

    sufficient to restore locomotor rhythmicity in SCN-lesioned

    hosts was demonstrated by encapsulating donor SCN tissue in a

    membrane that prevented neural outgrowth while allowing the

    diffusion of signals between graft and host (Silver et al., 1996).

    One candidate diffusible signal is prokineticin-2 (PK2)(Cheng et al., 2002). This protein is expressed rhythmically in

    the SCN, and its receptor is present in all major SCN targets

    (Cheng et al., 2002, 2005). Likewise, PK2 administration

    during the night (when levels are low) inhibits wheel running

    behavior. Whether or not this signal normally operates in a

    diffusible manner and/or is released synaptically requires

    further examination. A second candidate diffusible signal is

    transforming growth factor-alpha (TGF-alpha) (Kramer et al.,

    2001). As with PK2, TGF-alpha is expressed rhythmically in

    the SCN, and its administration inhibits wheel-running

    behavior. The receptor for TGF-alpha is also expressed in the

    SPVZ, the major target of the SCN. The degree which TGF-

    alpha is released in a diffusible manner under normal conditions

    is not known. Studies in which the contribution of neural

    efferents and diffusible signals can be distinguished are

    necessary to begin to answer this question.

    Although it is intriguing to speculate on the role of these

    signals in communicating information from the SCN, theproblem of unequivocally identifying an endogenous, physio-

    logically relevant diffusible SCN signal is complex and parallel

    in scope to the task faced by Sir Geoffrey Harris' in providing

    evidence for hypothalamic control over pituitary function in the

    1950s. The necessary and sufficient criteria to confirm the

    existence of a diffusible signal in a fluid volume have been

    summarized previously (Nicholson, 1999). First, the removal or

    replacement of the signaling substance must result in a change

    in the response being controlled. The substance should be

    present or increased, or both, in a well-defined temporal

    relationship to the response (and similarly declines when the

    response disappears). In addition, evidence must be obtained

    that a fluid compartment is the conduit for a diffusible or

    Fig. 2. A simplified model of the intracellular mechanisms responsible for mammalian circadian rhythm generation. The process begins when CLOCK and BMAL1

    proteins dimerize to drive the transcription of the Per(Per1, Per2, and Per3) and Cry (Cry1 and Cry2) genes. In turn,Perand Cry are translocated to the cytoplasm and

    translated into their respective proteins. Throughout the day, PER and CRY proteins rise within the cell cytoplasm. When levels of PER and CRY reach a threshold,

    they form heterodimers, feed back to the cell nucleus and negatively regulate CLOCK:BMAL1 mediated transcription of their own genes. This feedback loop takes

    approximately 24 h, thereby leading to an intracellular circadian rhythm.

    561L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    6/18

    562 L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    7/18

    transported signal. The signal must have access to and enter the

    compartment where the fluid dynamics and turnover in the

    compartment should allow appropriate movement of the signal.

    While PK2 and TGF-alpha meet some of these criteria, further

    research is necessary to clarify the role of these molecules in

    communicating circadian information.

    Neural control of neurosecretory factors

    In contrast to behavioral rhythms (e.g., locomotion, drinking,

    gnawing), endocrine rhythms require neural projections from

    the SCN to endocrine targets; endocrine rhythms are abolished

    after knife cuts severing SCN efferents (Hakim et al., 1991;

    Nunez and Stephan, 1977) and are not restored in SCN-lesioned

    transplanted animals (Meyer-Bernstein et al., 1999; Nunez and

    Stephan, 1977; Silver et al., 1996), presumably due to

    inadequate neural innervation of the host brain by the graft.

    Further evidence for a neural SCN output signal regulating

    hormone secretion is seen in studies of female hamsters. Whenhoused in constant light, the activity of a subset of hamsters

    splits into two separate activity bouts within a 24-h interval.

    These split females display two daily LH surges, each

    approximately half the concentration of a single surge in a

    non-split female (Swann and Turek, 1985) (Fig. 3). Under

    normal conditions, both halves of the bilaterally symmetrical

    SCN are active in synchrony. In ovariectomized, estrogen-

    implanted split hamsters killed during one of their activity

    bouts, however, activation of the SCN occurs on one side of the

    brain (monitored by FOS expression) but not on the other,

    suggesting that each half of the SCN can control an activity bout

    (de la Iglesia et al., 2000). Remarkably, FOS activation in

    GnRH neurons was only seen on the side of the brain in whichSCN FOS expression occurred (de la Iglesia et al., 2003). These

    findings suggest that the precise timing of the LH surge is

    derived from a neural signal originating in the SCN and

    communicated to ipsilateral GnRH neurons, as a diffusible

    output signal would reach both sides of the brain. Importantly,

    some hypothalamic sites are activated ipsilaterally, while others

    are activated either ipsilaterally or bilaterally in the split animal,

    again supporting the notion of multiple SCN output pathways

    (Tavakoli-Nezhad and Schwartz, 2005; Yan et al., 2005).

    Neural output from the SCN has been extensively

    investigated in rats and hamsters using tract tracing

    techniques (Kalsbeek et al., 1993; Kriegsfeld et al., 2004;Leak and Moore, 2001; Morin et al., 1994; Stephan et al.,

    1981; Watts and Swanson, 1987; Watts et al., 1987).

    Importantly, many of these monosynaptic projections target

    brain regions containing neuroendocrine cells producing

    hypothalamic releasing hormones (Fig. 4). Direct projections

    have been traced from the SCN to the medial preoptic area

    (MPOA), supraoptic nucleus (SON), anteroventral periven-

    tricular nucleus (AVPV), the paraventricular nucleus (PVN),

    the dorsomedial nucleus of the hypothalamus (DMH), and the

    lateral septum and the arcuate (Arc). The SCN also projects to

    the pineal through a multisynaptic pathway (Klein, 1985;

    Klein et al., 1983). There is abundant evidence for direct

    neural SCN control of neuroendocrine cell populations (Buijs

    et al., 1998, 2003; Egli et al., 2004; Gerhold et al., 2001;Horvath, 1997; Horvath et al., 1998; Kalsbeek and Buijs,

    2002; Kalsbeek et al., 1996a,b, 2000; Kriegsfeld et al., 2002a,

    b; Van der Beek et al., 1993, 1997b; Vrang et al., 1995).

    Because these cell populations can regulate neurochemicals

    that are secreted into the CSF (Reiter and Tan, 2002; Skinner

    and Caraty, 2002; Skinner and Malpaux, 1999; Tricoire et al.,

    2003) or general circulation, SCN-derived signals can control

    widespread systems in the brain and body.

    Together, the findings summarized above suggest several

    possibilities: behavioral rhythms may be controlled by a

    diffusible signal(s), while endocrine rhythms may require

    neural output. Alternatively, behavioral and endocrine rhythmscan both be supported by diffusible signals, but the threshold for

    supporting behavioral rhythms is lower. Finally, behavioral

    rhythms are controlled by both neural and diffusible signals, and

    either can maintain rhythmic function, while endocrine rhythms

    can only be supported via neural connections. Definitive

    identification of biologically significant endogenous diffusible

    signal(s) and the precise mode of SCN control is a current line

    of inquiry.

    Neural SCN output and estrus regulation

    SCN control of the rodent estrous cycle has been investigated

    extensively (Barbacka-Surowiak et al., 2003), providing anexcellent model system for investigations of circadian and

    neuroendocrine interactions. The SCN sends projections

    directly to GnRH neurons in female rodents (Horvath et al.,

    1998; Van der Beek et al., 1997a). These efferents express the

    SCN peptide, vasoactive intestinal polypeptide (VIP). GnRH

    neurons particularly important for the regulation of the estrous

    cycle are activated at the time of proestrus and receive SCN

    input (van der Beek et al., 1994). Also, sex differences in the

    daily expression of SCN VIP mRNA are seen in rats, with

    females exhibiting a peak 12 h out of phase with that of males

    (Krajnak et al., 1998). Presumably, the signal regulating the

    estrous cycle is sexually dimorphic, thereby lending furthersupport for VIP regulation of estrus. Furthermore, antisense

    oligonucleotides directed against VIP lead to a delayed and

    attenuated LH surge (Harney et al., 1996), reminiscent of that

    seen in middle-aged rats (Gore, 1998). Sex differences also exist

    in the pattern of projections from the SCN (Horvath et al., 1998;

    Van der Beek et al., 1997a,b). These sex differences in SCN

    projections upon GnRH cells, and in the production of

    Fig. 3. Circadian control of gonadotropin secretion. Syrian hamsters normally exhibit one consolidated bout of activity every 24 h. Under conditions of constant light,

    the hamsters activity splits intotwo components separated by about 12 h. In one ingenious study(de la Iglesia et al., 2003), the investigators killed animals prior to each

    of the activity bouts (see asterisk on activity records). Brains were analyzed for FOS activity in the SCN and in neurons of the GnRH neuronal system. Splitting

    behavior resulted in only one half of the SCN being active during a given time of day. GnRH was only activated (expressed FOS) on the side of the brain in which the

    SCN was active. Given that ovariectomized, estrogen-implanted hamsters with split behavior experience two LH surges (Swann and Turek, 1985), we are speculatingthat the neural mechanism underlying this phenomenon can be explain by differential leftright activation of the GnRH system at two times of day.

    563L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    8/18

    neurochemicals in SCN cells projecting to the GnRH system,

    may underlie the absence of an LH surge in males.In addition to direct connections to GnRH neurons, the SCN

    projects extensively to the anteroventral periventricular nucleus

    (AVPV), a brain region associated with the induction of the

    preovulatory LH surge (Le et al., 1997; Levine, 1997). The cells

    to which the SCN projects are estrogen-responsive (Watson et

    al., 1995), suggesting that the AVPV may be an important

    integration point for circadian and steroidal signals. Given the

    widespread projections of the SCN throughout the CNS, along

    with extensive input to the GnRH system, the potential for

    additional indirect modulation of the reproductive axis is

    considerable.

    Vasopressin, another SCN peptide important in the regula-

    tion of the estrous cycle, is synthesized and released in acircadian manner. SCN vasopressin is rhythmically secreted

    with a peak during a sensitive time window prior to the LH

    surge (Kalsbeek et al., 1996a,b). Vasopressin administration

    into the MPOA induces an LH surge in SCN-lesioned,

    ovariectomized rats treated with estradiol (Palm et al., 2001a,

    b). Electrical stimulation of the MPOA and VP administration

    into the MPOA induces an LH surge in SCN-lesioned rats

    (Palm et al., 1999). Finally, in co-cultures of POA and SCN

    tissue, the rhythm of GnRH release is in phase with the rhythm

    of VP release, but not with that of VIP (Funabashi et al., 2000).

    Paradoxically, Brattleboro rats incapable of synthesizing

    vasopressin are fertile, although abnormal estrous cyclicityand reduced fertility have been noted (Boer et al., 1982). Taken

    together, these data indicate a potential role for VP in inducing

    the LH surge, although it is likely not the sole mediator.

    Positive feedback effects of estrogen serve a permissive role

    in initiating the LH surge upon the arrival of the signal from the

    circadian pacemaker (Barbacka-Surowiak et al., 2003; Levine,

    1997); implants of an anti-estrogen into the POA block the LH

    surge (Petersen and Barraclough, 1989). This dependence on

    estrogen ensures the maturation of the follicle during the time of

    the surge, while the circadian dependence ensures that receptive

    behavior coincides with ovulation. It remains unclear, however,

    how these two signals converge at the cellular level to allow

    integration at the appropriate time of day. One possibility is that

    estrogen alters neurochemical secretion by the SCN or the

    signaling efficacy of these chemicals via second messengersystems and kinases (Chappell and Levine, 2000; Levine, 1997).

    An alternative hypothesis is that estrogen stimulates ligand-

    independent progesterone receptor production, and the timed

    neuronal signal acts on progesterone receptors (Chappell et al.,

    2000; Levine, 1997; Levine et al., 2001). This scheme suggests

    that the effects of estrogen are integrated with the SCN signal at

    the level of progesterone receptors to ensure that the GnRH

    system is sensitive to the daily signal only during the

    preovulatory estrogen surge. The fact that few estrogen receptors

    have been localized to GnRH neurons (Shivers et al., 1983)

    suggests that estrogen acts to produce progesterone in neurons

    upstream of the GnRH system and hints at important,

    unidentified, SCN projections to these upstream components.Given the potential importance of estrogen/progesterone

    receptor expressing neurons upstream of the GnRH system, it

    will be interesting to establish the means by which the circadian

    system communicates with and modulates these regulatory

    elements.

    The importance of a functional molecular clock in driving

    GnRH cells is seen in mice with a mutant form of the Clock

    gene. These mice have long, irregular estrous cycles and fail to

    exhibit an LH surge following estradiol treatment. Furthermore,

    this mutation also leads to an increase in fetal reabsorption

    during pregnancy and a decline in full-term parturition (Miller

    et al., 2004). These deficits are associated with a decline in mid-term levels of progesterone, suggesting abnormal secretion

    patterns of prolactin (Miller et al., 2004). Together, these

    findings highlight the importance of the circadian system in

    regulating the temporal pattern of hormone secretion necessary

    for mating, pregnancy, and its maintenance, although results

    using mutant models must be interpreted cautiously until

    converging approaches consistently support these conclusions.

    Not only does the SCN regulate GnRH during the estrous

    cycle, but other aspects of the estrous cycle are also regulated by

    the SCN. During proestrus, rising levels of estradiol reach a

    critical point and trigger the release of prolactin at a specific

    time of day. This release of prolactin is dependent upon the

    estradiol-induced increase in tuberoinfundibular dopaminergic

    Fig. 4. Efferent projections of the rodent SCN to its targets in the brain. Below each target area is a list of neuroendocrine cells that lie in that region of the brain and

    could potentially be regulated by direct projections from the SCN. Solid lines represent monosynaptic projections, while the dotted line represents a multisynaptic

    projection to the pineal gland. The pronounced overlap between neuroendocrine cells and SCN efferent terminals, combined with reports demonstrating direct

    neuronal projections from the SCN to neuroendocrine cells (e.g., GnRH and CRH cells), provides suggestive evidence for a global mechanism of circadian hormonal

    regulation. Adapted from Kriegsfeld et al. (2002a) with permission.

    564 L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    9/18

    (TIDA) neuron activity (Neill et al., 1971). Administration of an

    estradiol antiserum on the morning of diestrus-2 blocks the

    proestrous surge of prolactin (Neill et al., 1971). The proper

    timing of prolactin is achieved via SCN projections to TIDA

    neurons in the arcuate (Horvath, 1997). Additionally, TIDA

    neurons rhythmically express the clock gene, Per1, providing a

    potential additional means of temporal control (Kriegsfeld et al.,2003)and see below). SCN lesions result in an abolition of a

    daily prolactin rhythm (Mai et al., 1994) and abolish the

    preovulatory prolactin surge (Pan and Gala, 1985). Given the

    importance of prolactin timing in maintaining the corpus

    luteum, these findings further suggest an essential role for the

    circadian clock in reproduction and may explain the increased

    fetal reabsorption in Clockmutant mice described above (Miller

    et al., 2004).

    It has been well established that environmental factors act to

    fine-tune endogenous regulation of the reproductive cycle. For

    example, the timing of the prolactin surge is regulated by the

    phase of the light

    dark (LD) cycle. A change in the LD cycleleads to predictable changes in the timing of the estrogen-

    induced prolactin surge and the mating-induced prolactin surge

    (Blake, 1976; Pieper and Gala, 1979). Thus, environmental time

    of day information is transmitted to the circadian system to

    precisely coordinate reproductive events relative to local time.

    Cervically stimulated prolactin surges have a free-running cycle

    in ovariectomized, estradiol-treated rats held in constant

    conditions. This pattern is abolished after ablation of the SCN

    (Bethea and Neill, 1980). As with the LH surge, this finding

    suggests that both the timing and production of the prolactin

    surge require a signal from the circadian clock. The means by

    which environmental stimuli other than light (e.g., social signals,

    local conditions, nutrition, etc.) are integrated into this system tofine-tune the timing of hormonal and behavioral events are

    unknown and represent an opportunity for exploration.

    As with estrous behavior, the preovulatory LH surge

    occurs at regular 4- or 5-day intervals in rats, on the day of

    proestrus at a specific time of day coupled to the LD cycle

    (Colombo et al., 1974). Interestingly, despite the fact that the

    temporal pattern of SCN neural activity is similar in nocturnal

    and diurnal rodents (Smale et al., 2003), with a peak during

    the day and a trough at night, the timing of the preovulatory

    LH surge is reversed (Mahoney et al., 2004; McElhinny et al.,

    1999). Although the mechanisms by which this reversal

    occurs remain elusive, comparisons between diurnal andnocturnal species may provide insight into how circadian

    control is accomplished in humans (i.e., a diurnal species).

    Studies in rhesus initially indicated that the LH surge could

    be induced at any circadian phase in primates (Knobil, 1974).

    However, frequent urinary LH monitoring of women with

    regular menstrual cycles suggests a pronounced influence of

    circadian timing on the preovulatory LH surge, with most

    exhibiting the LH surge between midnight and 8:00 AM

    (Cahill et al., 1998; Edwards, 1981) (Fig. 5). Of 155 regular

    cycles monitored, 146 surges occurred during this 8-h time

    window (Cahill et al., 1998). Given that the daily timing of

    the LH surge in women can be unmasked under carefully

    monitored and controlled conditions, it is likely that the

    human ovulatory cycle also requires interactions between the

    circadian and reproductive systems.

    Direct and indirect transcriptional control as a clock output

    The circadian system exerts a widespread influence over

    numerous bodily functions. DNA microarray studies in mice

    indicate that59% of the genome, excluding genes involved

    in the core clock loop, are under circadian control (Akhtar et al.,

    2002; Panda et al., 2002; Storch et al., 2002). However, these

    so-called clock-controlled genes (CCGs) differ among tissues,

    with any two tissues likely sharing less than 10% of CCGsunder circadian control. Together, these findings suggest that

    circadian control is ubiquitous throughout the body, and tissue-

    specific processes may be controlled by differential activation

    of downstream genes in individual systems.

    Direct transcriptional control

    CCGs maintain a predictable phase relationship with the core

    clock genes (Ueda et al., 2002), indicating that the CCGs are

    either directly or indirectly regulated via the circadian

    transcriptional machinery. The expression of some CCGs is

    directly controlled by the CLOCK:BMAL1 heterodimer bindingto an E-box enhancer (CACGTG) in their promoter (Jin et al.,

    1999). An example of direct transcriptional control by the

    circadian system is seen with vasopressin regulation. Vasopres-

    sin is present in the SCN where it acts locally to regulate rhythm

    generation (Mihai et al., 1994a,b). It cannot be the sole

    regulatory factor, as the SCN of Brattleboro rats maintains

    rhythms in electrical activity (Ingram et al., 1998), and these

    animals exhibit only slight disruptions in rhythm amplitude and

    entrainability (Brown and Nunez, 1989; Murphy et al., 1993,

    1996). In addition to a role within the nucleus, SCN,

    vasopressin-expressing neurons signal distant hypothalamic

    targets and SCN vasopressin signaling has been implicated in

    the control of estrus (Buijs et al., 2003a,b; Kalsbeek et al., 1996a,

    Fig. 5. Frequency of onset of the LH surge by time of day. A total of 155 cycles

    from 35 women were monitored. The graph represents the percentage of

    preovulatory LH surges occurring during each time interval. Adapted from

    Cahill et al. (1998).

    565L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    10/18

    b; Mihai et al., 1994a,b; Palm et al., 2001a,b). The SCN rhythm

    in vasopressin (a rhythm that is not present in other

    vasopressinergic cell populationsee below) is dependent on

    an E-box element in the 5 flanking region of the vasopressin

    gene to which the CLOCK:BMAL complex binds. The

    vasopressin rhythm is abolished in mutant mice with aberrant

    Clock gene expression (Jin et al., 1999; Silver et al., 1999).E-box elements in the promoter have the potential for direct

    control by circadian clock genes, but this alone is not sufficient.

    In the SON (unlike the SCN), vasopressin is dependent on

    osmotic balance and is not rhythmic (Jin et al., 1999). While

    expression of Clock is robust in the SCN and SON, Bmal1 is

    expressed robustly only in the SCN and is barely detectable in

    the SON. These findings indicate some of the conditions

    necessary for direct control of genes regulating neuroendocrine

    function by the circadian clock transcriptional machinery.

    A gene with an E-box in its promoter region is a potential

    target for direct control by the transcriptional machinery of the

    circadian clock. We screened hypothalamic and pituitaryendocrine factors for E-box elements in the promoter of their

    published gene sequences to determine their potential for this

    means of control (Table 1). We found that some releasing

    hormones (TSH, GHRH, and vasopressin) do have E-box

    enhancers. We have shown in the adult mouse brain that a

    subset of neuroendocrine cells in the preoptic area, paraven-

    tricular nucleus of the hypothalamus, and the arcuate nucleus

    express the clock gene Per1 (Kriegsfeld et al., 2003), suggesting

    that this direct transcriptional regulation likely plays a key role

    in the organization of endocrine timing.

    Indirect transcriptional control

    In studies of the liver, it has been demonstrated that D-

    element binding protein (DBP) is another CCG under direct

    transcriptional regulation by the core circadian feedback loop

    (Ripperger et al., 2000). Importantly, DBP binds to other gene

    promoters to temporally regulate their transcription. This

    process is important in the control of hepatic metabolic

    processes; DBP activates the transcription of albumin, choles-

    terol 7 hydroxylase, and cytochrome P450 (Lavery et al.,

    1999). This second order system can provide ubiquitous control

    via the circadian system and temporal control of key enzymatic

    pathways involved in hormone production.

    A process similar to hepatic regulation by DBP may

    modulate the reproductive axis, as the gene for GnRH doesnot have an E-box but appears to be under circadian control. A

    series of studies using GT1 cells, an immortalized line of GnRH

    cells, demonstrated the rhythmic expression of numerous

    circadian clock genes (Chappell et al., 2003; Gillespie et al.,

    2003; Olcese et al., 2003). Importantly, GnRH release occurs

    episodically approximately every 90 min in most species, and

    GT1 cells also express this ultradian pattern of GnRH

    production. When clock genes are disrupted in GT1 cells, not

    only is the daily rhythm disrupted, but both pulse frequency and

    amplitude are also dramatically altered (Chappell et al., 2003).

    These findings suggest that, in addition to 24-h cycles, clock

    genes expressed within neuroendocrine cells may regulate theepisodic pattern of hormone secretion outside of the circadian

    range, even when the regulated neuroendocrine gene lacks an E-

    box enhancer. While these data are intriguing, a direct link

    between clock genes and ultradian rhythmicity awaits further

    supporting evidence.

    While the forgoing studies using embryonic cell lines

    provide insights into cellular mechanisms, how these data

    generalize to functioning in vivo needs to be determined, as

    immortalized cell lines may exhibit properties different from

    those of the living animal. In addition, cell lines lack the

    innumerable regulatory systems that act directly or indirectly on

    GnRH cells in vivo. To address the question of whether or not

    GnRH cells in adult mice express Per1 message, we used micewith a green fluorescent protein (GFP) reporter driven by Per1

    promoter. In these mice, GFP expression was observed in

    neurons near to GnRH-immunoreactive cells, but the two

    proteins were not co-localized. We have confirmed these

    negative findings using double-label immunofluroscence in

    mice and rats using several different antibodies against Per1 and

    GnRH (Kriegsfeld and Silver, unpublished data). We conclude

    that there are key differences between clock gene expression

    between cultured GT1 cells and GnRH neurons in the brains of

    adult animals.

    System-level control and coordination of endocrine function

    The task in understanding the orchestration of hormonal

    systems is best understood by recognizing that most processes

    in the body exhibit a circadian rhythm, and that activity in

    various systems exhibits different phases (i.e., peak and trough

    times) relative to each other. One explanation for how this feat is

    accomplished suggests that the SCN secretes one neurochem-

    ical to control each rhythmic process at its appropriate phase. In

    this view, the SCN secretes numerous substances, each

    precisely timed. Alternatively, control may be accomplished

    by the specificity of SCN projections and combinations of

    transmitter release at each target (Kalsbeek and Buijs, 2002). As

    an alternative hypothesis, we have suggested that SCN timing

    Table 1

    List of mammalian neuroendocrine factors containing an E-box (CACGTG)

    enhancer in their promoter

    Gene E-box (#) Reference

    Hypothalamic factors:CRH No (Muglia et al., 1994)

    GHRH Yes (1) (Laird, 2001direct submission)

    GnRH-I No (Hayflick et al., 1989)

    GnRH-II No (White et al., 1998)

    Ghrelin No (Kanamoto et al., 2004)

    Oxytocin No (Hara et al., 1990)

    Vasopressin Yes (1) (Hara et al., 1990; Jin et al., 1999)

    TRH No (Satoh et al., 1996)

    Pituitary:

    POMC (ACTH, MSH) No (Drouin et al., 1985)

    FSH No (Kumar, 1994direct submission)

    GH No (Das et al., 1996)

    LH No (Kaiser et al., 1998)

    Prolactin No (Gubbins et al., 1980)

    TSH (beta subunit) Yes (3) (Croyle and Maurer, 1984)

    566 L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    11/18

    signals have different consequences at each targeted effector

    system (Kriegsfeld et al., 2003).

    According to this view, a small number of rhythmic SCN

    signals must be differentially interpreted by a large number of

    targets to accomplish precise phase control by the SCN over an

    ensemble of rhythmic processes. Different targets respond

    differentially to the same message based on time of day (or localconditions), with some systems responding maximally to a

    signal(s) at a particular time of day, while another system might

    respond to this same signal(s) with inhibition. This hypothesis

    requires that target systems of the SCN have a mechanism for

    keeping time.

    Seasonal as well as circadian timing are dependent upon the

    SCN. Several hypotheses have been proposed to account for

    how the SCN and its targets track time on a seasonal basis (Carr

    et al., 2003; Hofman, 2004; Lincoln et al., 2003). In the SCN of

    Syrian and Siberian hamsters, photoperiod alters the duration of

    clock and clock-controlled gene expression, while the amplitude

    of gene expression is influenced by photoperiod in the parstuberalis (Johnston et al., 2003; Messager et al., 2000). In sheep,

    however, the relative timing of clock genes is altered by

    photoperiod in the pars tuberalis, providing a mechanism of

    temporal encoding and downstream control (Hazlerigg et al.,

    2004; Lincoln et al., 2002, 2003, 2005). These correlational

    results are intriguing and suggest that phase and/or amplitude of

    clock and CCGs in SCN brain targets and endocrine glands may

    predict their responsiveness to upstream signals on a daily

    schedule.

    Numerous systems display time-gated sensitivity to stimuli,

    with the same stimulus or neurochemical producing different

    effects at different times of day, suggesting temporal control at

    effector sites rather than passive regulation by the SCN. Thepreoptic area, for example, exhibits robust clock gene

    expression (Palm et al., 2001a,b; Tei et al., 1997; Yamamoto

    et al., 2001. Importantly, stimulation of the POA of ovariecto-

    mized rats with the SCN peptide, vasopressin, induces an LH

    surge during the second portion of the light period, but not the

    first (Palm, 2001 #122-001), indicating important temporal

    control at the level of the POA.

    Clocks in the neuroendocrine system

    The fact that time-keeping machinery is functional in

    numerous central neurosecretory cells (Kriegsfeld et al., 2003)and peripheral endocrine tissues (Bittman et al., 2003; Morse et

    al., 2003; Shieh, 2003; Von Gall et al., 2002; Zylka et al., 1998)

    represents a mechanism by which SCN targets can anticipate the

    reception of SCN signals and respond based upon local needs

    and time of day. In addition, because peripheral systems are

    controlled hierarchically by multiple upstream components,

    temporal modification in each link along the hypothalamo

    pituitaryendocrine gland axis could provide additional control

    over daily patterns of individual rhythms. At the top of this

    circadian hierarchy of control, the SCN sends signals to

    neuroendocrine cells and tissues to maintain synchronization

    among cellular oscillators in phenotypically distinct neuroen-

    docrine cell populations. Although capable of oscillating

    independently, these neuroendocrine cells respond to periodic

    SCN input to maintain a synchronized rhythm in the local cell

    population. Such mechanism could account for the loss of

    coherent circadian endocrine rhythms following lesions or

    transection of outputs from the SCN (Meyer-Bernstein et al.,

    1999; Moore and Eichler, 1972; Nunez and Stephan, 1977).

    Loss of coherence among elements would result in a bluntedoverall output and absence of rhythmicity when individual cells

    drift out of phase with each other.

    If the SCN coordinates cell populations, why do individual

    cells need their own clocks? According to this view, clocks in

    SCN target populations provide responsiveness to local

    conditions and temporal fine-tuning in a local population, not

    possible with a single driving master clock. These local clocks

    could selectively respond depending on time of day and

    appropriately drive the expression of cell/tissue-dependent

    CCGs that act as output to affect target systems or regulate

    local conditions. Coordinated rhythmic output from neuroen-

    docrine cells can then be communicated to the pituitary, whichalso exhibits circadian clock gene expression indicating a zone

    for further temporal modification (Messager et al., 2000).

    Finally, rhythmic information from the pituitary can be

    communicated humorally to target glands in the periphery that

    themselves express clock genes (Bittman et al., 2003; Zylka et

    al., 1998). As mentioned previously, multisynaptic projections

    from the SCN to several endocrine glands have been identified

    using viral tracers (Buijs et al., 1998, 1999; Gerendai and

    Halasz, 2000; Kalsbeek et al., 2000). These connections provide

    a mechanism for the SCN to coordinate peripheral cellular

    oscillators to optimize responses to slower endocrine signals.

    Several lines of evidence support the notion that neural

    communication from the SCN to the periphery is responsiblefor the timing of clock gene expression in targets organs and

    glands (Guo et al., 2005; Shibata, 2004; Terazono et al., 2003).

    The top of the hierarchy: neural SCN output

    An organization of this type requires that the SCN

    communicate (directly or indirectly) with neuroendocrine cells

    expressing clock genes. There is substantial evidence of direct

    projections from the SCN to neuroendocrine cells (Buijs et al.,

    1993; Horvath et al., 1998; Kriegsfeld et al., 2002a,b;

    Teclemariam-Mesbah et al., 1997; Van der Beek et al., 1997a,

    b; Vrang et al., 1995). Expression of the clock gene, Per1, isrhythmically expressed in the Arc (Abe et al., 2002) and

    exhibits a stress-induced increase in the PVH (Abe et al., 2002;

    Takahashi et al., 2001). Per1 has been localized to CRH-ir cells

    in the PVH (Takahashi et al., 2001), while the neurochemical

    phenotype of Per1-expressing cells in the Arc has been

    identified as dopaminergic providing a potential mechanism

    of control of prolactin rhythms and the preovulatory prolactin

    surge (Kriegsfeld et al., 2003).

    Hormonal and neural communication to glands

    Relative to the neural communication by the SCN to

    neuroendocrine cells, hormonal communication is slow.

    567L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    12/18

    However, by using the bloodstream as a route of communica-

    tion, hormones modulated by the circadian system can

    communicate rhythmic information throughout the body.

    Additional temporal control occurs at target glands and organs

    by using circadian clock machinery to modulate responsiveness

    to hormonal signals. If this means of temporal control is

    implemented at peripheral targets, necessary alterations in thetiming of organ/gland responsiveness due to changes in local

    conditions can be communicated rapidly to hormone-sensitive

    targets to adjust the timing of their circadian clocks. Given that

    numerous organs (e.g., liver, pancreas) and endocrine glands

    (e.g., testes, adipose tissue, adrenal gland) investigated to date

    receive autonomic innervation from the SCN (Bamshad et al.,

    1998; Bartness et al., 2001; Buijs et al., 1999, 2003; Kalsbeek et

    al., 2000; Olcese et al., 2003), these multisynaptic connections

    may provide a rapid means of clock resetting in peripheral

    tissues to ensure proper reception of slower diffusible

    communication (Fig. 6). A role of autonomic control in

    peripheral clock resetting comes from investigations in whichmanipulations of autonomic connections to liver reset clock

    gene expression in this organ (Terazono et al., 2003).

    In summary, global clock resetting may be accomplished via

    hormonal signals, as glucocorticoids can adjust the phase of

    peripheral circadian clock genes (Balsalobre et al., 2000).

    Whereas the role of hormones other than glucocorticoids in

    resetting peripheral oscillators has not been investigated, the

    marked effects of hormones on circadian function reviewed

    herein suggest a potentially crucial role for endocrine factors in

    orchestrating this multioscillatory arrangement.

    Conclusions and perspectives

    In general, we are not aware of the precision in the timing

    and coordination of numerous events in our bodies, unless it is

    disrupted (e.g., jet lag). However, processes as fundamental as

    the timing of sleep and its coordination with feelings of hunger

    are a manifestation of numerous physiological and biochemicalevents that change systematically and predictably over the

    course of the day. Given the numerous salient time cues in the

    environment, one might intuit that these daily changes are

    passive responses to environmental change. However, as

    reviewed here, daily rhythms are endogenously generated and

    are synchronized to external time cues in order to ensure that

    bodily processes are carried out at the appropriate, optimal time

    of day or night.

    Because most brain and bodily processes require a

    significant amount of time to achieve appropriate regulation

    the body must anticipate these changes and prepare

    accordingly in advance. For example, genomic actions ofsteroid hormones can take several hours to have their effects,

    and these hormones must be secreted prior to the time during

    which the behavior is best performed. Likewise, timed peak

    and trough hormone secretion may be required to prevent

    receptor down-regulation and desensitization. Because gener-

    ating new receptors requires significant metabolic energy and

    time, episodic hormone secretion may be required to allow for

    adequate receptor turnover. Thus, an endogenous time-

    keeping system is necessary to anticipate environmental

    change and initiate internal adjustments in advance of the

    Fig. 6. Overall organization of the circadian system. This organization is based on the postulation that rhythmic system physiology is controlled by a combination of

    neural and diffusible signals originating from the SCN. In this view, specific systems may be differentially regulated by SCN signals via local clocks allowing for morespecific responsiveness based upon local needs and time of day (see text for additional details).

    568 L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    13/18

    appropriate environmental time in order to coordinate

    innumerable bodily processes.

    The present overview shows how the circadian system

    controls the timing of hormone secretion using a number of

    mechanisms, including direct transcriptional and SCN neural

    control of neurosecretory factors, and control of glands by

    hormones, clock genes, and autonomic innervation. Becausehormones can have a widespread influence over physiology and

    behavior, and provide a means by which circadian information

    can be communicated systemically, it is important to determine

    how these rhythms are regulated. Not only are hormones

    modulated by the circadian system, but hormonal feedback to

    the SCN also influences circadian function (Dubocovich et al.,

    1996; Ellis and Turek, 1983; Hastings et al., 1997; Jechura et al.,

    2003; Labyak and Lee, 1995; Lewy and Sack, 1997; Morin et al.,

    1977). Together, these mechanisms controlling endocrine timing

    entail regulatory actions by the circadian system and provide

    extensive opportunities for empirical investigations of behav-

    iorally relevant systems.

    Acknowledgments

    We thank Dr. Lily Yan for the discussions and suggestions

    during the preparation of this review. We also thank Sean Duffy

    for the editorial and technical assistance. The work described in

    our laboratory was supported by NIH grants NS37919 (RS) and

    MH-12408 (LJK).

    References

    Abe,M., Herzog, E.D., Yamazaki, S., Straume, M., Tei, H., Sakaki, Y., Menaker,

    M., Block, G.D., 2002. Circadian rhythms in isolated brain regions.J. Neurosci. 22 (1), 350356.

    Abizaid, A., Mezei, G., Horvath, T.L., 2004. Estradiol enhances light-induced

    expression of transcription factors in the SCN. Brain Res. 1010 (12),

    3544.

    Akhtar, R.A., Reddy, A.B., Maywood, E.S., Clayton, J.D., King, V.M., Smith,

    A.G., Gant, T.W., Hastings, M.H., Kyriacou, C.P., 2002. Circadian cycling

    of the mouse liver transcriptome, as revealed by cDNA microarray, is driven

    by the suprachiasmatic nucleus. Curr. Biol. 12 (7), 540550.

    Albrecht, U., 2004. The mammalian circadian clock: a network of gene

    expression. Front. Biosci. 9, 4855.

    Andrews, W.W., Ojeda, S.R., 1981. A detailed analysis of the serum luteinizing

    hormone secretory profile in conscious, free-moving female rats during the

    time of puberty. Endocrinology 109 (6), 20322039.

    Ashmore, L.J., Sehgal, A., 2003. A fly's eye view of circadian entrainment.

    J. Biol. Rhythms 18 (3), 206216.Balsalobre, A., Damiola, F., Schibler, U., 1998. A serum shock induces

    circadian gene expression in mammalian tissue culture cells. Cell 93 (6),

    929937.

    Balsalobre, A., Brown, S.A., Marcacci, L., Tronche, F., Kellendonk, C.,

    Reichardt, H.M., Schutz, G., Schibler, U., 2000. Resetting of circadian time

    in peripheral tissues by glucocorticoid signaling. Science 289 (5488),

    23442347.

    Bamshad, M., Aoki, V.T., Adkison, M.G., Warren, W.S., Bartness, T.J.,

    1998. Central nervous system origins of the sympathetic nervous system

    outflow to white adipose tissue. Am. J. Physiol. 275 (1 Pt. 2),

    R291R299.

    Barbacka-Surowiak, G., Surowiak, J., Stoklosowa, S., 2003. The involvement of

    suprachiasmatic nuclei in the regulation of estrous cycles in rodents. Reprod.

    Biol. 3 (2), 99129.

    Bartness, T.J., Song, C.K., Demas, G.E., 2001. SCN efferents to peripheral

    tissues: implications for biological rhythms. J. Biol. Rhythms 16 (3),

    196204.

    Bethea, C.L., Neill, J.D., 1980. Lesions of the suprachiasmatic nuclei abolish the

    cervically stimulated prolactin surges in the rat. Endocrinology 107 (1), 15.

    Bittman, E.L., Doherty, L., Huang, L., Paroskie, A., 2003. Period gene

    expression in mouse endocrine tissues. Am. J. Physiol.: Regul., Integr.

    Comp. Physiol. 285 (3), R561R569.

    Blake, C.A., 1976. Effects of intravenous infusion of catecholamines on ratplasma luteinizing hormone and prolactin concentrations. Endocrinology 98

    (1), 99104.

    Blaustein, J.D., Tetel, M.J., Ricciardi, K.H., Delville, Y., Turcotte, J.C., 1994.

    Hypothalamic ovarian steroid hormone-sensitive neurons involved in female

    sexual behavior. Psychoneuroendocrinology 19 (57), 505516.

    Boer, G.J., Boer, K., Swaab, D.F., 1982. On the reproductive and

    developmental differences within the Brattleboro strain. Ann. N. Y.

    Acad. Sci. 394, 3745.

    Boyar, R.M., Wu, R.H., Roffwarg, H., Kapen, S., Weitzman, E.D., Hellman, L.,

    Finkelstein, J.W., 1976. Human puberty: 24-hour estradiol in pubertal girls.

    J. Clin. Endocrinol. Metab. 43 (6), 14181421.

    Brown, M.H., Nunez, A.A., 1989. Vasopressin-deficient rats show a reduced

    amplitude of the circadian sleep rhythm. Physiol. Behav. 46 (4), 759762.

    Buijs, R.M., Markman, M., Nunes-Cardoso, B., Hou, Y.X., Shinn, S., 1993.

    Projections of the suprachiasmatic nucleus to stress-related areas in the rathypothalamus: a light and electron microscopic study. J. Comp. Neurol. 335

    (1), 4254.

    Buijs, R.M., Hermes, M.H., Kalsbeek, A., 1998. The suprachiasmatic nucleus-

    paraventricular nucleus interactions: a bridge to the neuroendocrine and

    autonomic nervous system. Prog. Brain Res. 119, 365382.

    Buijs, R.M., Wortel, J., Van Heerikhuize, J.J., Feenstra, M.G., Ter Horst, G.J.,

    Romijn, H.J., Kalsbeek, A., 1999. Anatomical and functional demonstration

    of a multisynaptic suprachiasmatic nucleus adrenal (cortex) pathway. Eur. J.

    Neurosci. 11 (5), 15351544.

    Buijs, R.M., la Fleur, S.E., Wortel, J., Van Heyningen, C., Zuiddam, L.,

    Mettenleiter, T.C., Kalsbeek, A., Nagai, K., Niijima, A., 2003a. The

    suprachiasmatic nucleus balances sympathetic and parasympathetic output

    to peripheral organs through separate preautonomic neurons. J. Comp.

    Neurol. 464 (1), 3648.

    Buijs, R.M., van Eden, C.G., Goncharuk, V.D., Kalsbeek, A., 2003b. Thebiological clock tunes the organs of the body: timing by hormones and the

    autonomic nervous system. J. Endocrinol. 177 (1), 1726.

    Cahill, D.J., Wardle, P.G., Harlow, C.R., Hull, M.G., 1998. Onset of the

    preovulatory luteinizing hormone surge: diurnal timing and critical follicular

    prerequisites. Fertil. Steril. 70 (1), 5659.

    Carr, A.J., Johnston, J.D., Semikhodskii, A.G., Nolan, T., Cagampang, F.R.,

    Stirland, J.A., Loudon, A.S., 2003. Photoperiod differentially regulates

    circadian oscillators in central and peripheral tissues of the Syrian hamster.

    Curr. Biol. 13 (17), 15431548.

    Chappell, P.E., Levine, J.E., 2000. Stimulation of gonadotropin-releasing

    hormone surges by estrogen: I. Role of hypothalamic progesterone

    receptors. Endocrinology 141 (4), 14771485.

    Chappell, P.E., Lee, J., Levine, J.E., 2000. Stimulation of gonadotropin-

    releasing hormone surges by estrogen: II. Role of cyclic adenosine 35-

    monophosphate. Endocrinology 141 (4), 14861492.Chappell, P.E., White, R.S., Mellon, P.L., 2003. Circadian gene expression

    regulates pulsatile gonadotropin-releasing hormone (GnRH) secretory

    patterns in the hypothalamic GnRH-secreting GT1-7 cell line. J. Neurosci.

    23 (35), 1120211213.

    Cheng, M.Y., Bullock, C.M., Li, C., Lee, A.G., Bermak, J.C., Belluzzi, J.,

    Weaver, D.R., Leslie, F.M., Zhou, Q.Y., 2002. Prokineticin 2 transmits the

    behavioural circadian rhythm of the suprachiasmatic nucleus. Nature 417

    (6887), 405410.

    Cheng, M.Y., Bittman, E.L., Hattar, S., Zhou, Q.Y., 2005. Regulation of

    prokineticin 2 expression by light and the circadian clock. BMC Neurosci. 6

    (1), 17.

    Clancy, A.N., Whitman, C., Michael, R.P., Albers, H.E., 1994. Distribution of

    androgen receptor-like immunoreactivity in the brains of intact and castrated

    male hamsters. Brain Res. Bull. 33 (3), 325332.

    Colombo, J.A., Baldwin, D.M., Sawyer, C.H., 1974. Timing of the estrogen-

    569L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    14/18

    induced release of LH in ovariectomized rats under an altered lighting

    schedule. Proc. Soc. Exp. Biol. Med. 145 (3), 11251127.

    Croyle, M.L., Maurer, R.A., 1984. Thyroid hormone decreases thyrotropin

    subunit mRNA levels in rat anterior pituitary. DNA 3 (3), 231236.

    Das, P., Meyer, L., Seyfert, H.M., Brockmann, G., Schwerin, M., 1996.

    Structure of the growth hormone-encoding gene and its promoter in mice.

    Gene 169 (2), 209213.

    de la Iglesia, H.O., Blaustein, J.D., Bittman, E.L., 1999. Oestrogen receptor-alpha-immunoreactive neurones project to the suprachiasmatic nucleus of

    the female Syrian hamster. J. Neuroendocrinol. 11 (7), 481490.

    de la Iglesia, H.O., Meyer, J., Carpino Jr., A., Schwartz, W.J., 2000. Antiphase

    oscillation of the left and right suprachiasmatic nuclei. Science 290 (5492),

    799801.

    de la Iglesia, H.O., Meyer, J., Schwartz, W.J., 2003. Lateralization of circadian

    pacemaker output: activation of left- and right-sided luteinizing hormone-

    releasing hormone neurons involves a neural rather than a humoral pathway.

    J. Neurosci. 23 (19), 74127414.

    Desir, D., Van Cauter, E., L'Hermite, M., Refetoff, S., Jadot, C., Caufriez, A.,

    Copinschi, G., Robyn, C., 1982. Effects of jet lag on hormonal patterns:

    III. Demonstration of an intrinsic circadian rhythmicity in plasma prolactin.

    J. Clin. Endocrinol. Metab. 55 (5), 849857.

    Drouin, J., Chamberland, M., Charron, J., Jeannotte, L., Nemer, M., 1985.

    Structure of the rat pro-opiomelanocortin (POMC) gene. FEBS Lett. 193 (1),5458.

    Du, Y.Z., Tong, J., 2002. Genetic regulation of circadian clock. Sheng Li Ke Xue

    Jin Zhan 33 (4), 343345.

    Dubocovich, M.L., Benloucif, S., Masana, M.I., 1996. Melatonin receptors in

    the mammalian suprachiasmatic nucleus. Behav. Brain Res. 73 (12),

    141147.

    Duffield, G.E., 2003. DNA microarray analyses of circadian timing: the

    genomic basis of biological time. J. Neuroendocrinol. 15 (10), 9911002.

    Edwards, R.G., 1981. Test-tube babies, 1981. Nature 293 (5830), 253256.

    Egli, M., Bertram, R., Sellix, M.T., Freeman, M.E., 2004. Rhythmic secretion

    of prolactin in rats: action of oxytocin coordinated by vasoactive

    intestinal polypeptide of suprachiasmatic nucleus origin. Endocrinology

    33863394.

    Ellis, G.B., Turek, F.W., 1983. Testosterone and photoperiod interact to regulate

    locomotor activity in male hamsters. Horm. Behav. 17 (1), 6675.Everett, J.W., Sawyer, C.H., 1950. A 24-hour periodicity in the LH-

    release apparatus of female rat disclosed by barbiturate administration.

    Endocrinology 47, 198218.

    Fernandez-Guasti, A., Kruijver, F.P., Fodor, M., Swaab, D.F., 2000. Sex

    differences in the distribution of androgen receptors in the human

    hypothalamus. J. Comp. Neurol. 425 (3), 422435.

    Funabashi, T., Shinohara, K., Mitsushima, D., Kimura, F., 2000. Gonadotropin-

    releasing hormone exhibits circadian rhythm in phase with arginine-

    vasopressin in co-cultures of the female rat preoptic area and suprachias-

    matic nucleus. J. Neuroendocrinol. 12 (6), 521528.

    Gachon, F., Nagoshi, E., Brown, S.A., Ripperger, J., Schibler, U., 2004. The

    mammalian circadian timing system: from gene expression to physiology.

    Chromosoma 103112.

    Gerendai, I., Halasz, B., 2000. Central nervous system structures connected with

    the endocrine glands. Findings obtained with the viral transneuronal tracingtechnique. Exp. Clin. Endocrinol. Diabetes 108 (6), 389395.

    Gerhold, L.M., Horvath, T.L., Freeman, M.E., 2001. Vasoactive intestinal

    peptide fibers innervate neuroendocrine dopaminergic neurons. Brain Res.

    919 (1), 4856.

    Gillespie, J.M., Chan, B.P., Roy, D., Cai, F., Belsham, D.D., 2003. Expressionof

    circadian rhythm genes in gonadotropin-releasing hormone-secreting GT1-7

    neurons. Endocrinology 144 (12), 52855292.

    Glossop, N.R., Hardin, P.E., 2002. Central and peripheral circadian

    oscillator mechanisms in flies and mammals. J. Cell Sci. 115

    (Pt. 17), 33693377.

    Gore, A.C., 1998. Circadian rhythms during aging. In: Mobbs, C.V., Hof,

    P.R. (Eds.), Functional Endocrinology of Aging, vol. 29. Karger,

    Basel, pp. 127165.

    Green, C.B., 2003. Molecular control of Xenopus retinal circadian rhythms.

    J. Neuroendocrinol. 15 (4), 350354.

    Green, D.J., Gillette, R., 1982. Circadian rhythm of firing rate recorded from

    single cells in the rat suprachiasmatic brain slice. Brain Res. 245 (1),

    198200.

    Groos, G., Hendriks, J., 1982. Circadian rhythms in electrical discharge of rat

    suprachiasmatic neurones recorded in vitro. Neurosci. Lett. 34 (3),

    283288.

    Gubbins, E.J., Maurer, R.A., Lagrimini, M., Erwin, C.R., Donelson, J.E., 1980.

    Structure of the rat prolactin gene. J. Biol. Chem. 255 (18), 86558662.Gundlah, C., Kohama, S.G., Mirkes, S.J., Garyfallou, V.T., Urbanski, H.F.,

    Bethea, C.L., 2000. Distribution of estrogen receptor beta (ERbeta) mRNA

    in hypothalamus, midbrain and temporal lobe of spayed macaque: continued

    expression with hormone replacement. Brain Res. Mol. Brain Res. 76 (2),

    191204.

    Guo, H., Brewer, J.M., Champhekar, A., Harris, R.B., Bittman, E.L., 2005.

    Differential control of peripheral circadian rhythms by suprachiasmatic-

    dependent neural signals. Proc. Natl. Acad. Sci. U. S. A. 102 (8),

    31113116.

    Hakim, H., DeBernardo, A.P., Silver, R., 1991. Circadian locomotor rhythms,

    but not photoperiodic responses, survive surgical isolation of the SCN in

    hamsters. J. Biol. Rhythms 6 (2), 97113.

    Hansen, S., Sodersten, P., Eneroth, P., Srebro, B., Hole, K., 1979. A sexually

    dimorphic rhythm in oestradiol-activated lordosis behaviour in the rat.

    J. Endocrinol. 83 (2), 267274.Hara, Y., Battey, J., Gainer, H., 1990. Structure of mouse vasopressin and

    oxytocin genes. Brain Res. Mol. Brain Res. 8 (4), 319324.

    Harney, J.P., Scarbrough, K., Rosewell, K.L., Wise, P.M., 1996. In vivo

    antisense antagonism of vasoactive intestinal peptide in the suprachiasmatic

    nuclei causes aging-like changes in the estradiol-induced luteinizing

    hormone and prolactin surges. Endocrinology 137 (9), 36963701.

    Hastings, M.H., 1991. Neuroendocrine rhythms. In: Redfern, P.H., Waterhouse,

    J. (Eds.), Pharmacological Therapy, vol. 50. Pergamon Press, Great Britain,

    pp. 3571.

    Hastings, M.H., Duffield, G.E., Ebling, F.J., Kidd, A., Maywood, E.S., Schurov,

    I., 1997. Non-photic signalling in the suprachiasmatic nucleus. Biol. Cell 89

    (8), 495503.

    Hastings, M.H., Reddy, A.B., Maywood, E.S., 2003. A clockwork web:

    circadian timing in brain and periphery, in health and disease. Nat. Rev.,

    Neurosci. 4 (8), 649661.Hayflick, J.S., Adelman, J.P., Seeburg, P.H., 1989. The complete nucleotide

    sequence of the human gonadotropin-releasing hormone gene. Nucleic

    Acids Res. 17 (15), 64036404.

    Hazlerigg, D.G., Andersson, H., Johnston, J.D., Lincoln, G., 2004. Molecular

    characterization of the long-day response in the Soay sheep, a seasonal

    mammal. Curr. Biol. 14 (4), 334339.

    Hofman, M.A., 2004. The brain's calendar: neural mechanisms of seasonal

    timing. Biol. Rev. Camb. Philos. Soc. 79 (1), 6177.

    Hofman, M.A., Fliers, E., Goudsmit, E., Swaab, D.F., 1988. Morphometric

    analysis of the suprachiasmatic and paraventricular nuclei in the human

    brain: sex differences and age-dependent changes. J. Anat. 160, 127143.

    Hofman, M.A., Zhou, J.N., Swaab, D.F., 1996. Suprachiasmatic nucleus of the

    human brain: an immunocytochemical and morphometric analysis. Anat.

    Rec. 244 (4), 552562.

    Horvath, T.L., 1997. Suprachiasmatic efferents avoid phenestrated capillariesbut innervate neuroendocrine cells, including those producing dopamine.

    Endocrinology 138 (3), 13121320.

    Horvath, T.L., Cela, V., van der Beek, E.M., 1998. Gender-specific apposition

    between vasoactive intestinal peptide-containing axons and gonadotrophin-

    releasing hormone-producing neurons in the rat. Brain Res. 795 (12),

    277281.

    Ingram, C.D., Ciobanu, R., Coculescu, I.L., Tanasescu, R., Coculescu, M.,

    Mihai, R., 1998. Vasopressin neurotransmission and the control of circadian

    rhythms in the suprachiasmatic nucleus. Prog. Brain Res. 119, 351364.

    Jakacki, R.I., Kelch, R.P., Sauder, S.E., Lloyd, J.S., Hopwood, N.J., Marshall,

    J.C., 1982. Pulsatile secretion of luteinizing hormone in children. J. Clin.

    Endocrinol. Metab. 55 (3), 453458.

    Jechura, T.J., Walsh, J.M., Lee, T.M., 2003. Testosterone suppresses circadian

    responsiveness to social cues in the diurnal rodent Octodon degus. J. Biol.

    Rhythms 18 (1), 4350.

    570 L.J. Kriegsfeld, R. Silver / Hormones and Behavior 49 (2006) 557574

  • 7/31/2019 154 kriegsfeld et al

    15/18

    Jin, X., Shearman, L.P., Weaver, D.R., Zylka, M.J., de Vries, G.J., Reppert,

    S.M., 1999. A molecular mechanism regulating rhythmic output from the

    suprachiasmatic circadian clock. Cell 96 (1), 5768.

    Johnston, J.D., Cagampang, F.R., Stirland, J.A., Carr, A.J., White, M.R., Davis,

    J.R., Loudon, A.S., 2003. Evidence for an endogenous per1- and ICER-

    independent seasonal timer in the hamster pituitary gland. FASEB J. 17 (8),

    810815.

    Kaiser, U.B., Sabbagh, E., Chen, M.T., Chin, W.W., Saunders, B.D., 1998. Sp1binds to the rat luteinizing hormone beta (LHbeta) gene promoter and

    mediates gonadotropin-releasing hormone-stimulated expression of the

    LHbeta subunit gene. J. Biol. Chem. 273 (21), 1294312951.

    Kalsbeek, A., Buijs, R.M., 2002. Output pathways of the mammalian

    suprachiasmatic nucleus: coding circadian time by transmitter selection

    and specific targeting. Cell Tissue Res. 309 (1), 109118.

    Kalsbeek, A., Teclemariam-Mesbah, R., Pevet, P., 1993. Efferent projections of

    the suprachiasmatic nucleus in the golden hamster (Mesocricetus auratus).

    J. Comp. Neurol. 332 (3), 293314.

    Kalsbeek, A., van der Vliet, J., Buijs, R.M., 1996a. Decrease of endogenous

    vasopressin release necessary for expression of the circadian rise in plasma

    corticosterone: a reverse microdialysis study. J. Neuroendocrinol. 8 (4),

    299307.

    Kalsbeek, A., van Heerikhuize, J.J., Wortel, J., Buijs, R.M., 1996b. A diurnal

    rhythm of stimulatory input to the hypothalamo-pituitary-adrenal system asrevealed by timed intrahypothalamic administration of the vasopressin V1

    antagonist. J. Neurosci. 16 (17), 55555565.

    Kalsbeek, A., Fliers, E., Franke, A.N., Wortel, J., Buijs, R.M., 2000. Functional

    connections between the suprachiasmatic nucleus and the thyroid gland as

    revealed by lesioning and viral tracing techniques in the rat. Endocrinology

    141 (10), 38323841.

    Kanamoto, N., Akamizu, T., Tagami, T., Hataya, Y., Moriyama, K., Takaya, K.,

    Hosoda, H., Kojima, M., Kangawa, K., Nakao, K., 2004. Genomic structure

    and characterization of the 5-flanking region of the human ghrelin gene.

    Endocrinology 41444153.

    Kapen, S., Boyar, R.M., Finkelstein, J.W., Hellman, L., Weitzman, E.D., 1974.

    Effect of sleepwake cycle reversal on luteinizing hormone secretory pattern

    in puberty. J. Clin. Endocrinol. Metab. 39 (2), 293299.

    Kashon, M.L., Arbogast, J.A., Sisk, C.L., 1996. Distribution and hormonal

    regulation of androgen receptor immunoreactivity in the forebrain of themale European ferret. J. Comp. Neurol. 376 (4), 567586.

    Klein, D.C., 1985. Photoneural regulation of the mammalian pineal gland. Ciba

    Found. Symp. 117, 3856.

    Klein, D.C., Smoot, R., Weller, J.L., Higa, S., Markey, S.P., Creed, G.J.,

    Jacobowitz, D.M., 1983. Lesions of the paraventricular nucleus area of

    the hypothalamus disrupt the suprachiasmatic leads to spinal cord circuit

    in the melatonin rhythm generating system. Brain Res. Bull. 10 (5),

    647652.

    Knobil, E., 1974. On the control of gonadotropin secretion in the rhesus

    monkey. Recent Prog. Horm. Res. 30 (0), 146.

    Krajnak, K., Kashon, M.L., Rosewell, K.L., Wise, P.M., 1998. Sex differences

    in the daily rhythm of vasoactive intestinal polypeptide but not arginine

    vasopressin messenger ribonucleic acid in the suprachiasmatic nuclei.

    Endocrinology 139 (10), 41894196.

    Kramer, A., Yang, F.C., Snodgrass, P., Li, X., Scammell, T.E., Davis, F.C.,Weitz, C.J., 2001. Regulation of daily locomotor activity and sleep by

    hypothalamic EGF receptor signaling. Science 294 (5551), 25112515.

    Kriegsfeld, L.J., LeSauter, J.L., Hamada, T., Pitts, S.M., Silver, R., 2002a.

    Circadian rhythms in the endocrine system. In: Pfaff, D., Etgen, A. (Eds.),

    Hormones, Brain, and Behavior. Academic Press, New York.

    Kriegsfeld, L.J., Silver, R., Gore, A.C., Crews, D., 2002b. Vasoactive intestinal

    polypeptide contacts on gonadotropin-releasing hormone neurones increase

    following puberty in female rats. J. Neuroendocrinol. 14 (9), 685690.

    Kriegsfeld, L.J., Korets, R., Silver, R., 2003. Expression of the circadian clock

    gene Period 1 in neuroendocrine cells: an investigation using mice with a

    Per1:GFP transgene. Eur. J. Neurosci. 17 (2), 212220.

    Kriegsfeld, L.J., Leak, R.K., Yackulic, C.B., LeSauter, J., Silver, R., 2004.

    Organization of suprachiasmatic nucleus projections in Syrian hamsters

    (Mesocricetus auratus): an anterograde and retrograde analysis. J. Comp.

    Neurol. 468 (3), 361379.

    Kruijver, F.P., Swaab, D.F., 2002. Sex hormone receptors are present in the

    human suprachiasmatic nucleus. Neuroendocrinology 75 (5), 296305.

    Labyak, S.E., Lee, T.M., 1995. Estrus- and steroid-induced changes in circadian

    rhythms in a diurnal rodent, Octodon degus. Physiol. Behav. 58 (3),

    573585.

    Lavery, D.J., Lopez-Molina, L., Margueron, R., Fleury-Olela, F., Conquet, F.,

    Schibler, U., Bonfils, C., 1999. Circadian expression of the steroid 15 alpha-

    hydroxylase (Cyp2a4) and coumarin 7-hydroxylase (Cyp2a5) genes inmouse liver is regulated by the PAR leucine zipper transcription factor DBP.

    Mol. Cell. Biol. 19 (10), 64886499.

    Le, W.W., Attardi, B., Berghorn, K.A., Blaustein, J., Hoffman, G.E., 1997.

    Progesterone blockade of a luteinizing hormone surge blocks luteinizing

    hormone-releasing hormone Fos activationand activation of its preoptic area

    afferents. Brain Res. 778 (2), 272280.

    Leak, R.K., Moore, R.Y., 2001. Topographic organization of suprachiasmatic

    nucleus projection neurons. J. Comp. Neurol. 433 (3), 312334.

    Lehman, M.N., Silver, R., Gla